Skip to main content
Log in

Pairing correlations within the micro-macroscopic approach for the level density

  • Regular Article - Theoretical Physics
  • Published:
The European Physical Journal A Aims and scope Submit manuscript

Abstract

Level density \(\rho (E,N,Z)\) is calculated for the two-component close- and open-shell nuclei with a given energy E, and neutron N and proton Z numbers, taking into account pairing effects within the microscopic-macroscopic approach (MMA). These analytical calculations have been carried out by using semiclassical statistical mean-field approximations beyond the saddle-point method of the Fermi gas model in a low excitation-energies range. The level density \(\rho \), obtained as function of the system entropy S, depends essentially on the condensation energy \(E_{\textrm{cond}}\) through the excitation energy U in super-fluid nuclei. The simplest super-fluid approach, based on the BCS theory, accounts for a smooth temperature dependence of the pairing gap \(\Delta \) due to particle number fluctuations. Taking into account the pairing effects in magic or semi-magic nuclei, excited below neutron resonances, one finds a notable pairing phase transition. Pairing correlations sometimes improve significantly the comparison with experimental data.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Data availability statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: The calculated values of K were used for the restricted number of nuclei in the specific problem, and do not have a systematic character, and therefore, they are not required to be deposited.]

References

  1. T. Ericson, Adv. Phys. 9, 425 (1960)

    CAS  ADS  Google Scholar 

  2. A.A. Bohr, B.R. Mottelson, Nuclear Structure, vol. 1 (Benjamin, New York, 1967)

    Google Scholar 

  3. L.D. Landau, E.M. Lifshitz, Statistical Physics, Course of Theoretical Physics, vol. 5 (Pergamon, Oxford, 1975)

    Google Scholar 

  4. A.V. Ignatyuk, Statistical Properties of Excited Atomic Nuclei (Energoatomizdat, 1983). (Russian)

    Google Scholar 

  5. N.K. Grossjean, H. Feldmeier, Nucl. Phys. A 444, 113 (1985)

    ADS  Google Scholar 

  6. Yu.V. Sokolov, Level Density of Atomic Nuclei (Energoatomizadat, 1990). (Russian)

    Google Scholar 

  7. S. Shlomo, Nucl. Phys. A 539, 17 (1992)

    ADS  Google Scholar 

  8. A.V. Ignatyuk, Level densities, in Handbook for Calculations of Nuclear Reaction Data. (International Atomic Energy Agency, Vienna, 1998), pp.65–80

    Google Scholar 

  9. Y. Alhassid, G.F. Bertsch, L. Fang, Phys. Rev. C 68, 044322 (2003)

    ADS  Google Scholar 

  10. T. von Egidy, D. Bucurescu, Phys. Rev. C 78, 051301(R) (2008)

    ADS  Google Scholar 

  11. T. von Egidy, D. Bucurescu, Phys. Rev. C 80, 054310 (2009)

    ADS  Google Scholar 

  12. Y. Alhassid, G.F. Bertsch, C.N. Gilbreth, H. Nakada, Phys. Rev. C 93, 044320 (2016)

    ADS  Google Scholar 

  13. V.M. Kolomietz, A.I. Sanzhur, S. Shlomo, Phys. Rev. C 97, 064302 (2018)

    CAS  ADS  Google Scholar 

  14. V. Zelevinsky, A. Volya, Mesoscopic Nuclear Physics: From Nucleus To Quantum Chaos To Quantum Signal (World Scientific, Singapore, 2018)

    Google Scholar 

  15. V. Zelevinsky, M. Horoi, Prog. Part. Nucl. Phys. 105, 180 (2019)

    CAS  ADS  Google Scholar 

  16. V.M. Kolomietz, S. Shlomo, Mean Field Theory (World Scientific, Cham, 2020)

    Google Scholar 

  17. R. Bengtsson, P. Schuck, Phys. Lett. B 89, 321 (1980)

    ADS  Google Scholar 

  18. P. Ring, P. Schuck, The Nuclear Many-Body Problem (Springer-Verlag, New York, 1980)

    Google Scholar 

  19. J.L. Egido, P. Ring, S. Iwasaki, H.J. Mang, Phys. Lett. B 154, 1 (1985)

    ADS  Google Scholar 

  20. H. Kucharek, P. Ring, P. Schuck, R. Bengtsson, M. Girod, Phys. Lett. B 216, 249 (1989)

    CAS  ADS  Google Scholar 

  21. H. Kucharek, P. Ring, P. Schuck, Z. Phys. A 334, 119 (1989)

    CAS  ADS  Google Scholar 

  22. M. Farine, P. Schuck, X. Viñas, Phys. Rev. A 62, 013608 (2000)

    ADS  Google Scholar 

  23. X. Viñas, P. Schuck, M. Farine, M. Centelles, Phys. Rev. C 67, 054314 (2003)

    ADS  Google Scholar 

  24. M. Baldo, C. Maieron, P. Schuck, X. Viñas, Nucl. Phys. A 736, 241 (2004)

    ADS  Google Scholar 

  25. F. Barranco, P.F. Bortignon, R.A. Broglia, G. Colo, P. Schuck, E. Vigezzi, X. Viñas, Phys. Rev. C 72, 054314 (2005)

    ADS  Google Scholar 

  26. X. Viñas, P. Schuck, M. Farine, arXiv:1010.5961v1 [nucl.phys] (2010)

  27. X. Viñas, P. Schuck, M. Farine, J. Phys: Conf. Ser. 321, 012024 (2011)

    Google Scholar 

  28. A. Pastore, J. Margueron, P. Schuck, X. Viñas, Phys. Rev. C 88, 034314 (2013)

    ADS  Google Scholar 

  29. P. Schuck, X. Viñas, Fifty Years of Nuclear BCS (World Scientific, 2013), pp.212–226

    Google Scholar 

  30. A. Pastore, P. Schuck, X. Viñas, J. Margueron, Phys. Rev. A 90, 043634 (2014)

    ADS  Google Scholar 

  31. A. Pastore, P. Schuck, X. Viñas, arXiv:2004.09423 [nucl-th] (2020)

  32. P. Schuck, M. Urban, X. Viñas, Eur. Phys. J. A 59, 164 (2023)

    CAS  ADS  Google Scholar 

  33. A.I. Levon et al., Phys. Rev. C 102, 014308 (2020)

    CAS  ADS  Google Scholar 

  34. V.M. Kolomietz, A.G. Magner, V.M. Strutinsky, Sov. J. Nucl. Phys. 29, 758 (1979)

    Google Scholar 

  35. A.G. Magner, A.I. Sanzhur, S.N. Fedotkin, A.I. Levon, S. Shlomo, Int. J. Mod. Phys. E 30, 2150092 (2021)

    CAS  ADS  Google Scholar 

  36. A.G. Magner, A.I. Sanzhur, S.N. Fedotkin, A.I. Levon, S. Shlomo, Phys. Rev. C 104, 044319 (2021)

    CAS  ADS  Google Scholar 

  37. A.G. Magner, A.I. Sanzhur, S.N. Fedotkin, A.I. Levon, S. Shlomo, Nucl. Phys. A 1021, 122423 (2022)

    CAS  Google Scholar 

  38. A.G. Magner, A.I. Sanzhur, S.N. Fedotkin, A.I. Levon, U.V. Grygoriev, S. Shlomo, Low Temp. Phys. 48, 920 (2022)

    CAS  ADS  Google Scholar 

  39. A.G. Magner, A.I. Sanzhur, S.N. Fedotkin, A.I. Levon, U.V. Grygoriev, S. Shlomo, Nucl. Phys. At. Energy 24, 175 (2023)

    ADS  Google Scholar 

  40. M. Brack, L. Damgaard, A.S. Jensen, H.C. Pauli, V.M. Strutinsky, C.Y. Wong, Rev. Mod. Phys. 44, 320 (1972)

    CAS  ADS  Google Scholar 

  41. V.M. Strutinsky, On the nuclear level density in case of an energy gap, in Proceedings of the International Conference on Nuclear Physics. (Springer, 1958), pp.617–622

    Google Scholar 

  42. A. Bohr, B.R. Mottelson, D. Pines, Phys. Rev. 110, 936 (1958)

    CAS  ADS  Google Scholar 

  43. S.T. Belyaev, Mat. Fys. Medd. Dan. Vid. Selsk 31, 3 (1959)

    Google Scholar 

  44. J. Bardeen, L.N. Cooper, J.R. Schrieffer, Phys. Rev. 108, 1175 (1957)

    MathSciNet  CAS  ADS  Google Scholar 

  45. L.D. Landau, E.M. Lifshitz, Statistical Physics, Part 2, Course of Theoretical Physics, vol. 9 (Pergamon, London, 1980)

    Google Scholar 

  46. A. Sedrakian, J.W. Clark, Eur. Phys. J. A 55, 167 (2019)

    CAS  ADS  Google Scholar 

  47. B. Le Crom et al., Phys. Lett. B 829, 137057 (2022)

    Google Scholar 

  48. V.M. Strutinsky, A.G. Magner, Sov. J. Part. Nucl. 7, 138 (1976)

    Google Scholar 

  49. V.M. Strutinsky, A.G. Magner, S.R. Ofengenden, T. Døssing, Z. Phys. A 283, 269 (1977)

    ADS  Google Scholar 

  50. M. Brack, R.K. Bhaduri, Semiclassical Physics Frontiers in Physics, No. 96, vol. 2 (Westview Press, Boulder, 2003)

  51. A.G. Magner, Y.S. Yatsyshyn, K. Arita, M. Brack, Phys. At. Nucl. 74, 1445 (2011)

    CAS  Google Scholar 

  52. M. Brack, C. Guet, H.-B. Håkansson, Phys. Rep. 123, 275 (1985)

    CAS  ADS  Google Scholar 

  53. P. Moeller, A.J. Sierk, T. Ichikawa, H. Sagawa, At. Data Nucl. Data Tables 109–110, 1–204 (2016)

    ADS  Google Scholar 

  54. P. Vogel, B. Jonson, P.G. Hansen, Phys. Lett. 139, 227 (1984)

    Google Scholar 

  55. National Nuclear Data Center On-Line Data Service for the ENSDF database, http://www.nndc.bnl.gov/ensdf

Download references

Acknowledgements

We dedicate this article to the memory of Professor Peter Schuck. The authors gratefully acknowledge A. Bonasera, M. Brack, D. Bucurescu, A.N. Gorbachenko, E. Koshchiy, S.P. Maydanyuk, J.B. Natowitz, E. Pollacco, V.A. Plujko, P. Ring, A. Volya, and S. Yennello for many discussions, constructive suggestions, and fruitful help. This work was partially supported by the project No. 0120U102221 of the National Academy of Sciences of Ukraine. S. Shlomo and A.G. Magner are partially supported by the US Department of Energy under Grant no. DE-FG03-93ER-40773.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. G. Magner.

Additional information

Communicated by David Blaschke

Appendix A: Semiclassical periodic orbit theory

Appendix A: Semiclassical periodic orbit theory

For the sake of a simple notation, we shall presently restrict ourselves to the case of A nucleons (one kind only) in a given local (HF) potential \(V(\textbf{r})\) as in Refs. [35, 48, 50,51,52]. The level density shell corrections can be presented analytically within the periodic orbit theory (POT) in terms of the sum over classical periodic orbits (PO) [48, 50, 51],

$$\begin{aligned}{} & {} {\displaystyle \delta g_{\textrm{scl}}(\varepsilon )= \sum ^{}_{\textrm{PO}}g^{}_{\textrm{PO}}(\varepsilon ),}\nonumber \\{} & {} {\displaystyle g^{}_{\textrm{PO}}(\varepsilon )\!=\!{\mathcal {A}}_\textrm{PO}(\varepsilon ) \,\cos \left[ \frac{1}{\hbar }{\mathbb {S}}_\textrm{PO}(\varepsilon )- \frac{\pi }{2} \mu ^{}_{\textrm{PO}} -\phi ^{}_0\right] \!.} \end{aligned}$$
(A1)

Here \({\mathbb {S}}_{\textrm{PO}}(\varepsilon )\) is the classical action along the PO in the nucleon potential well of the same radius, \(R=r_0A^{1/3}\) (\(r^{}_0\approx 1.14\) fm), \(\mu ^{}_{\textrm{PO}}\) is the so called Maslov index, determined by the catastrophe points (turning and caustic points) along the PO, and \(\phi ^{}_0\) is an additional shift of the phase coming from the dimension of the problem and degeneracy of the POs. The amplitude \({\mathcal {A}}_\textrm{PO}(\varepsilon )\), and the action \({\mathbb {S}}_\textrm{PO}(\varepsilon )\), are smooth functions of the energy \(\varepsilon \). In addition, the amplitude, \({\mathcal {A}}_{\textrm{PO}}(\varepsilon )\), depends on the PO stability factors. The Gaussian local averaging of the level density shell correction, \(\delta g_\textrm{scl}(\varepsilon )\), over the single-particle (s.p.) energy spectrum \(\varepsilon ^{}_i\) near the Fermi surface \(\varepsilon ^{}_F\), with a width parameter \(\gamma \), smaller than a distance between major shells, \({\mathcal {D}}_{\textrm{sh}}\), can be done analytically [48, 50, 51],

$$\begin{aligned} \delta g^{}_{\gamma \mathrm scl}(\varepsilon ) \cong \sum ^{}_\textrm{PO}g^{}_{\textrm{PO}}(\varepsilon )~ \exp \left[ -\left( \frac{\gamma t^{}_{\textrm{PO}}}{2\hbar }\right) ^2\right] ~, \end{aligned}$$
(A2)

where \(t^{}_{\textrm{PO}}= \partial {\mathbb {S}}_{\textrm{PO}}/\partial \varepsilon \) is the period of particle motion along the PO in the potential well.

The smooth ground-state energy of the nucleus is approximated by \({\tilde{E}}\approx E_{\mathrm{\texttt{ETF}}}=\int _0^{\lambda } \hbox {d}\varepsilon ~\varepsilon ~ {\tilde{g}}(\varepsilon )\) , where \({\tilde{g}}(\varepsilon )\) is a smooth level density equal approximately to the ETF level density, \({\tilde{g}}\approx g_\mathrm{\texttt{ETF}}\), (\(\lambda \approx {\tilde{\lambda }}\), and \({\tilde{\lambda }}\) is the smooth chemical potential in the SCM). The chemical potential \(\lambda \) (or \(\tilde{\lambda }\)) is the solution of the corresponding conservation of particle number equation:

$$\begin{aligned} A = \int \limits _0^{\lambda }\text{ d } \varepsilon ~g(\varepsilon ) ~. \qquad \end{aligned}$$
(A3)

The POT shell component of the free energy, \(\delta F_{\textrm{scl}}\), is related in the non-thermal and non-rotational limit to the shell correction energy of a cold nucleus, \(\delta E_{\textrm{scl}}\), see Refs. [48, 50, 51]. Within the POT, \(\delta E_{\textrm{scl}}\) is determined, in turn, by the oscillating level density \(\delta g_\textrm{scl}(\varepsilon )\), Eq. (A1),

$$\begin{aligned} \delta E_{\textrm{scl}} \approx \sum _{\textrm{PO}}\frac{\hbar ^2}{t_\textrm{PO}^2}\delta g^{}_{\textrm{PO}}(\lambda )~. \end{aligned}$$
(A4)

The chemical potential \(\lambda \) can be approximated by the Fermi energy \(\varepsilon ^{}_F\), up to a small excitation energy, and isotopic asymmetry corrections (\(T\ll \lambda \) for the saddle point value \(T=1/\beta ^*\), if exists). It is determined by the particle-number conservation conditions, Eq.  (A3), where \(g(\varepsilon )\cong g_{\textrm{scl}}= g^{}_{\mathrm{\texttt{ETF}}} +\delta g_{\textrm{scl}}\) is the total POT level density. One now needs to solve equation (A3) to determine the chemical potential \(\lambda \) as function of the particle number A since \(\lambda \) is needed in Eq. (A4) to obtain the semiclassical energy shell correction \(\delta E_{\textrm{scl}}\).

For a major shell structure near the Fermi energy surface \(\varepsilon \approx \lambda \), the POT energy shell correction, \(\delta E_{\textrm{scl}}\), is approximately proportional to the level density shell correction, \(\delta g_{\textrm{scl}}(\varepsilon )\) [Eq. (A1)], at \(\varepsilon =\lambda \), Eq. (A4),

$$\begin{aligned} \delta E \approx \delta E_{\textrm{scl}}\approx \left( \frac{{\mathcal {D}}_{\textrm{sh}}}{2 \pi }\right) ^2~ \delta g_\textrm{scl}(\lambda )~, \end{aligned}$$
(A5)

where \({\mathcal {D}}_{\textrm{sh}} \approx \lambda /A^{1/3}\) is the mean distance between major nuclear shells. Indeed, the rapid convergence of the PO sum in Eq. (A4) is guaranteed by the factor in front of the density component \(g^{}_{\textrm{PO}}\), Eq. (A1), a factor which is inversely proportional to the period time \(t_{\textrm{PO}}(\lambda )\) squared along the PO. Therefore, only POs with short periods which occupy a significant phase-space volume near the Fermi surface will contribute. These orbits are responsible for the major shell structure, that is related to a Gaussian averaging width, \(\gamma \approx \gamma _{\textrm{sh}}\), which is much larger than the distance between neighboring s.p. states but much smaller than the distance \({\mathcal {D}}_{\textrm{sh}} \) between major shells near the Fermi surface. Eq. (A2) for the averaged s.p. level density was derived under these conditions for \(\gamma \). According to the POT [48, 50, 51], the distance between major shells, \({\mathcal {D}}_{\textrm{sh}}\), is determined by a mean period of the shortest and most degenerate POs, \(\langle t^{}_\textrm{PO}\rangle \) [50]:

$$\begin{aligned} {\mathcal {D}}_{\textrm{sh}} \cong \frac{2\pi \hbar }{\langle t_\textrm{PO}\rangle } \approx \frac{\lambda }{A^{1/3}}~. \end{aligned}$$
(A6)

Taking the factor in front of \(g^{}_{\textrm{PO}}\), in Eq. (A4) off the sum over the POs for the energy shell correction \(\delta E_{\textrm{scl}}\), one arrives at its semiclassical expression (A5) [48, 49, 51]. Differentiating Eq. (A5) with (A1) with respect to \(\lambda \) and keeping only the dominating terms coming from differentiation of the sine of the action phase argument, \({\mathbb {S}}_{\textrm{PO}}/\hbar \sim A^{1/3}\), one finds the useful relationships:

$$\begin{aligned}{} & {} {\displaystyle \frac{\partial ^2\delta E_\textrm{PO}}{\partial \lambda ^2} \approx -\delta g^{}_{\textrm{PO}}(\lambda )~},\nonumber \\{} & {} {\displaystyle \frac{\partial ^2 g_\textrm{scl}}{\partial \lambda ^2}\approx \sum ^{}_\textrm{PO}\frac{\partial ^2\delta g^{}_{\textrm{PO}}}{\partial \lambda ^2} \approx -\left( \frac{2\pi }{D_{\textrm{sh}}}\right) ^2 \delta g^{}_\textrm{PO}(\lambda )~}. \end{aligned}$$
(A7)

Notice that taking into account Eq. (A3) for the chemical potential \(\lambda \), one has another useful relationship for the second derivative of background thermodynamic potential, \(\Omega _0=\int _0^\lambda \hbox {d}\varepsilon (\varepsilon -\lambda ) g_\textrm{scl}(\varepsilon )\):

$$\begin{aligned} \frac{\partial ^2\Omega _{0}}{\partial \lambda ^2} \approx - g^{}_\textrm{scl}(\lambda )~. \end{aligned}$$
(A8)

The level density parameter a [see Eq. (10) for the entropy S] can be related to the averaged POT level density ETF and shell correction components by

$$\begin{aligned} a \approx a^{}_{\textrm{ETF}}+\delta a_{\textrm{scl}}= \frac{\pi ^2}{6} (g^{}_{\textrm{ETF}}+\delta g_{\textrm{scl}})~, \end{aligned}$$
(A9)

For shell corrections, one has the following relation:

$$\begin{aligned} \delta a_{\textrm{scl}}=\frac{\pi ^2}{6}\delta g_{\textrm{scl}}(\lambda ). \end{aligned}$$
(A10)

Using Eqs. (A9), (A10), and Eq. (A7), one obtains Eq. (9).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Magner, A.G., Sanzhur, A.I., Fedotkin, S.N. et al. Pairing correlations within the micro-macroscopic approach for the level density. Eur. Phys. J. A 60, 6 (2024). https://doi.org/10.1140/epja/s10050-023-01222-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epja/s10050-023-01222-1

Navigation